Vous êtes sur la page 1sur 9

ARTICLE

Received 7 Dec 2011 | Accepted 8 May 2012 | Published 12 Jun 2012

DOI: 10.1038/ncomms1891

Direct printing of nanostructures by electrostatic


autofocussing of ink nanodroplets
P. Galliker1, J. Schneider1, H. Eghlidi1,2, S. Kress1, V. Sandoghdar2, & D. Poulikakos1

Nanotechnology, with its broad impact on societally relevant applications, relies heavily on the
availability of accessible nanofabrication methods. Even though a host of such techniques exists,
the flexible, inexpensive, on-demand and scalable fabrication of functional nanostructures
remains largely elusive. Here we present a method involving nanoscale electrohydrodynamic
ink-jet printing that may significantly contribute in this direction. A combination of nanoscopic
placement precision, soft-landing fluid dynamics, rapid solvent vapourization, and subsequent
self-assembly of the ink colloidal content leads to the formation of scaffolds with base diameters
equal to that of a single ejected nanodroplet. The virtually material-independent growth of
nanostructures into the third dimension is then governed by an autofocussing phenomenon
caused by local electrostatic field enhancement, resulting in large aspect ratio. We demonstrate
the capabilities of our electrohydrodynamic printing technique with several examples, including
the fabrication of plasmonic nanoantennas with features sizes down to 50nm.

1 Laboratory of Thermodynamics in Emerging Technologies, Institute of Energy Technology, Department of Mechanical and Process Engineering,

ETH Zrich, CH-8092 Zrich, Switzerland. 2 Laboratory of Physical Chemistry, ETH Zrich, CH-8092 Zrich, Switzerland. Present address:
Max Planck Institute for the Science of Light, 91058 Erlangen, Germany. Correspondence and requests for materials should be addressed to
D.P. (email: dimos.poulikakos@ethz.ch).
nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

nature communications | DOI: 10.1038/ncomms1891

large pallet of techniques has been explored in the past few


decades for nanofabrication, including powerful approaches
such as photo- and electron-beam lithography, focussed-ion
beam milling as well as nanoparticle-based self-assembly17. A particularly attractive approach is printing by ink jet4, electrohydrodynamics1,5,7,8 and related techniques2,3, in which a nanoparticleladen ink is dispensed from a scanning nozzle. Although these
strategies provide flexible and inexpensive fabrication platforms,
various physical and experimental processes have limited their resolution. Electrohydrodynamic (EHD) liquid ejection, well known
from its use in electro-spinning9 and -spraying10, has recently been
employed in microfabrication for the production of defined planar
entities of the order of a micrometre1,7.
EHD ejection of liquid from a pipette nozzle is induced by the
relaxation of ions to the liquid surface, on activation by an electric
field. The ions are naturally present in the liquid (same amount of
positive and negative species) and are its electrically conductive
species. The induced electric stresses at the liquidair interface can
ultimately overcome the resisting effect of surface tension, causing
liquid ejection. Depending on the direction of the electric field,
either an excess of positive or negative ions will drive the ejection of
liquid and attribute to its non-zero electric charge. Whereas different EHD ejection modes have been identified11, most related applications, including those related to microfabrication1,7, are based on
the so-called cone-jet ejection mode in which a thin jet is released
from the tip of a larger sharply converging liquid cone at high volume flow rates1012. Another mode of ejection, called microdripping5,11,12, results in the periodic ejection of single, micron-sized
spherical droplets from the apex of a larger steady liquid meniscus.
Here we downscale this mode to the regime of nanodripping,
while avoiding clogging of the dispenser tip by a sufficient reduction
of the nanoparticle concentration. We will show that after droplet
impact and solvent vapourization, nanoparticles are uniformly distributed in an area essentially equivalent to the diameter of a single
droplet. In addition, we take advantage of an electrostatic nanodroplet autofocussing (ENA) effect to generate high aspect ratio out
of plane nanostructures, built from the nanoparticles contained
in the individual nanodroplets. Working together, the capabilities
of this process allow the realization of nanostructures of different
geometries and materials, which have homogeneous lateral size
equal to that of a single nanodroplet.

Results
Working principle and set-up. Figure 1a displays the essential
steps in the temporal sequence of ENA NanoDrip printing. In the
first step, a voltage is applied between the pipette/ink and a counter
electrode, resulting in the generation of a meniscus and periodic
ejection of ink nanodroplets from its lowest point. Once ejected,
the nanodroplet is accelerated by the electric field towards the
substrate where viscous effects pronounced by the nanodroplet
size facilitate a soft landing and, owing to wetting-enhanced rapid
vapourization, the droplet can dry during an ejection period e.
Once the solvent has vapourized, the nanoparticles, formerly
dispensed in the ink droplet, are found to be uniformly spread in an
area corresponding to the projection of the spherical nanodroplet
(Fig. 1b). Periodic occurrence of these events leads to a narrow
accumulation of nanoparticles on the substrate. As this structure
grows in height (Fig. 1c,d), its tip acts as a sharp electrode, creating
strong electric-field gradients that focus the incoming droplets
towards it and allow the structure to grow at a homogeneous
diameter. The interval between voltage initiation and termination is
denoted as pulse, as it has a defined duration and a sharp rise and fall
(Fig. 1a). However, the pulse length does not influence the ejection
frequency, nor the ejection process in general. The only purpose
of using pulses is to precisely control the number of deposited
droplets, that is, nanoparticles, and therewith the extent of printed


nanostructures. It is worth stressing that our terminology of a pulse


should not be confused with that of pulsed EHD jet-type processes
reported by other authors7,8. In these studies, the pulse duration is
related to the amount of liquid accumulated on the substrate, and
is therefore an important process parameter affecting the lateral
size of printed structures, after evaporation of the accumulated
liquid. In contrast, the remarkable features of the nanodripping
mode attained here, allow the solvent of each ejected and deposited
droplet to be removed before the impact of the next one. The extent
of accumulated liquid is therefore independent of the duration of
the applied pulse.
The experimental set-up is shown schematically in Fig.1e and
discussed in more detail in the Methods. The essential requirements
and related physical mechanisms needed to achieve the nanodripping mode are also commented on in the Methods. If not specifically mentioned, the ink in all experiments consisted of a dispersion
of 37nm gold nanoparticles in n-tetradecane. The nozzle was fixed
at a separation of 34m from the substrate, and a piezo-stage was
used to position the substrate with nanometer accuracy at will. The
printed structures were detected and imaged during deposition by
an in-house-built microscope (Fig. 1e and Methods). We remark,
however, that the nanodroplets could not be detected during flight
because of their large velocity magnitude of up to the order of
100 ms1 (Supplementary Methods and Supplementary Fig. S1).
We used scanning electron microscopy (SEM) for a quantitative
analysis of the nanoparticle patterns of a single droplet after deposition and solvent vapourization (Fig. 1b). For brevity, these patterns
will be denoted as footprints in the following. Footprints are of
crucial importance to ENA NanoDrip printing, because they allow
the determination of the droplet size and ejection frequency. In addition, footprints help us understand the physical mechanisms behind
the emergence of the resulting nanostructure geometries. Single
footprints were obtained by fixing the nozzle in space while separating the sequential droplet impact positions through a sufficiently
rapid movement of the substrate by the piezoelectric stage. With
the knowledge of the substrate velocity and the spatial separation of
footprints on the substrate, it is possible to evaluate an exact value
for the ejection frequency (Fig. 2a). The lateral extent of footprints
was found to approximately represent the droplet diameter. This
unexpected result is based on an analysis of the volume of printed
nanostructures (Supplementary Fig. S2) in relation to the applied
pulse length. Dividing the resulting ejection flow rate (ejected
volume per pulse length) by the experimentally obtained ejection
frequency yielded droplet sizes approximately matching the size
of footprints (Supplementary Fig. S3 and Supplementary Methods).
It is also found that the largest footprint diameters match the outer
diameter of the employed pipette. As no other forces than the electric and surface tension forces are present (gravitation is negligible),
the largest possible droplets are indeed expected to be similarly sized
as the outer pipette diameter. Further reasoning that strengthens the
equality argument of droplet and footprint sizes is provided later.
Taking into account that footprints approximately represent the
size of an ejected spherical droplet, their spatial extent could be
employed to determine the size of droplets as a function of applied
voltage (Fig. 2a). Combining frequency and size, the ejection flow
rate could be determined (Fig. 2b). We find that with a nozzle of
outer diameter ~1,200nm, footprints as small as 80nm were produced at an ejection frequency of several tens of kilohertz. To illustrate the homogeneous nature of frequency and droplet size in the
nanodripping mode, we show in Supplementary Fig. S4 a series of
separated footprints. These footprints are practically equidistant
and of equal size. These and all other experiments were performed
without active feedback control of the nozzlesubstrate distance.
That active feedback control is not a necessary requirement for a
reproducible printing process is also confirmed by additional considerations and experiments on the sensitivity of the results on the

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

Applied voltage

nature communications | DOI: 10.1038/ncomms1891

Time

Printing stage

0V

XYZ micro-stage
Micropipette
Substrate

Counter electrode

XYZ piezo-stage

Microscope

Control unit

Electroactuation
HV amplifier

iSCAT
400 Vp

Figure 1 | Schematic of ENA printing process and set-up. (a) Growth of a liquid meniscus and subsequent ejection of ink nanodroplets from its apex on
application of a DC voltage. During DC on-time, droplets are ejected at a homogeneous period e and, once impacted, are vapourized (represented by
wavy arrows) in the course of e. After periodic repetition of this event (for illustration convenience merged into one cycle), a sharp structure consisting of
a multitude of formerly dispersed nanoparticles rises from the substrate, attracting approaching charged droplets by ENA (straight arrows). The growth
process is further illustrated with SEM micrographs (150nm scale bar) of (b) the deposition pattern of a single nanodroplet and that of (c,d) actual
nanopillars. (e) Schematic of the ENA NanoDrip set-up with the nozzle located above an underlying glass substrate placed on an ITO-coated glass slide
representing the grounded counter electrode. Voltage stimuli were applied in the form of amplified DC signals between the ink-filled, metal-coated pipette
and the counter electrode. A homebuilt iSCAT imaging system was used to detect printed structures.

distance between nozzle and substrate (Supplementary Fig. S5 and


Supplementary Methods).
Numerical and experimental analysis of the ejection process. To
estimate the diameter of the ejected droplets theoretically, we use a
semi-numerical approach where we employ a simple geometrical
model in which the small droplet to be ejected is pendant at the
lower end of a hemispherical meniscus (Fig. 2c). On the basis of
this model, we then calculate the forces acting onto the small droplet. Although the hemispherical meniscus shape has been observed
before5,12,13, we experimentally verified that it develops with a
similar shape also for the small nozzle sizes employed in our study
(Supplementary Fig. S6). Along this line, the size of the droplets can be predicted by a balance of electric and surface tension
forces, whereby the electric force is calculated by the finite element
method (Fig. 2c shows the resulting electric field distribution) with
the assumption that nozzle, meniscus and pendant droplet are all

equipotential (see Methods). Green symbols in Fig. 2a show the


calculated droplet diameters as a function of the applied voltages.
The agreement with the experimental data is very good up to the
regime marked by the dashed arrow, where the liquid charge relaxation time (ratio of liquid permittivity and electrical conductivity)
equals the ejection period, e, rendering the equipotential assumption invalid. We find that this regime change also corresponds to a
minimum in the flow rate. For larger applied voltages, intensified
electric stresses are produced at the meniscus, whereas charges cannot relax to the pendant droplet surface fast enough. As a result,
a large liquid flow cannot be counteracted by a further reduction
of the pendant droplet diameter, and the flow rate strongly increases
for e below ~. That we can reproduce both the evolution of the
droplet diameter for e> and the divergence of experimental and
theoretical data for e< is proof for the appropriateness of our
simple semi-numerical model (within its limited applicability). It
must be mentioned that the purpose of the theoretical estimate

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

10
1
100

0.1
100

150

200
250
300
Applied voltage (V)

b
100

10

1
100

150

200

250

300

350

Applied voltage (V)

0 V m1
5e7
1e8
1.5e8
2e8
2.5e8
Z

3e8

Figure 2 | Summary of measured and numerically derived key quantities


for a 1.2m nozzle. (a) Experimentally obtained footprint diameter (full
black squares) and ejection frequency (open circles) and numerically
derived droplet diameter (full green triangles connected by straight green
lines) as a function of applied voltage. The dashed arrow indicates the
voltage threshold at which the charge relaxation time () equals the
ejection period (e). Data points are based on up to 50 averaged values
with standard errors below 5% (diameter) and 10% (frequency). The
highest errors reach 10% (diameter) and 30% (frequency). (b) Flow rate
of ejected fluid calculated from experimental footprint and frequency data.
(c) Nozzle geometry used for numerical calculation of the electric force
acting onto the liquid. The surface plot shows the calculated z-component
of the electric field around the modelled nozzle region during droplet
ejection. The nozzle is depicted with a green line and the meniscus with
a blue line. The image depicts a state in which the surface tension of
the pendant droplet is matched by the electric stress and the droplet is
therefore ready to be ejected.

is not to reproduce the ejection dynamics (flow rate estimates are


not possible). This task is highly complex and deserves a separate
dedicated study, beyond the purpose of the present experimental
work. Here we are only interested in a simple prediction of the size
of ejected droplets, as direct observation was not possible.

300
200
Applied voltage (V)

100

Nondimensional droplet
diameter

Flow rate (m3 s1)

100

350



1,000

100

10

1
0.1
0.1
1.0

1.5
2.0
2.5
Nondimensional voltage

Ejection frequency (kHz)

100

Diameter (nm)

1,000

Ejection frequency (kHz)

Diameter (nm)

nature communications | DOI: 10.1038/ncomms1891

3.0

Figure 3 | Reproducibility with respect to different nozzle diameters.


(a) Experimental (black symbols) and numerically (green symbols)
derived footprint/droplet diameter for a 1.2-m nozzle (squares, reprinted
from Fig. 2) and a 600-nm nozzle (triangles). The minimal ejection
voltage for the smaller nozzle was found to be ~20V lower than that of
the larger one. (b) The same experimental footprint data (squares) after
nondimensionalizing (droplet size divided by nozzle size, applied voltage
divided by minimal ejection voltage), with full symbols representing
the large and open symbols representing the small nozzle. The fact that
the datasets merge suggests that footprint sizes do indeed represent
the droplet size, which is elucidated by additional experiments in the
Supplementary Methods. We also plot the ejection frequency in absolute
values (circles). At the same nondimensional voltage, frequencies
observed for both nozzles, are about the same. This suggests an almost
constant ejection frequency independent of the nozzle diameter.

In Fig. 3a, we show, in addition to the footprint sizes of Fig. 2a,


a second dataset where a nozzle with a diameter of 600nm (half of
that used before) was employed. It is found that the minimal ejection voltage (the voltage needed for ejecting a droplet with a diameter equal to that of the meniscus) decreases with decreasing nozzle
diameter. In the Supplementary Methods, we illustrate theoretically
the generality of this behaviour. Also plotted in Fig. 3a are numerical estimates of the droplet size. Evidently, the footprint diameters
agree well with the theoretical estimates of the droplet size, confirming reproducibility of the numerical results. The presented data were
then nondimensionalized by dividing the actual footprint diameter
by the diameter of the meniscus, and by dividing the applied voltage
by the minimal ejection voltage (see Supplementary Methods for
reasoning). Figure 3b shows that the diameters of the nondimensional footprints are practically identical. In both cases, the smallest
droplet size is ~1/15 of the nozzle size. The merging of the dimensionless datasets as well as the agreement of our numerical estimate
with the experimental results substantially underpins the claimed
equality of footprint and droplet diameter. In Fig. 3a, we also plot
the dimensional ejection frequency as a function of nondimensional voltage. The ejection frequency seems to be nearly unaffected
by the nozzle size variation, which implies that the flow rate at the
same nondimensional voltage approximately scales with the droplet

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

nature communications | DOI: 10.1038/ncomms1891

0 Q0

1 Q0

2 Q0

Figure 4 | Electrostatic field-focussing effect for charged and uncharged nanopillars. The images show nanopillars subjected to an applied electric
field. The arrows point to the direction of acceleration of a charged nanodroplet with an arbitrary intensity represented by the arrow length. The figure
compares the resulting field for three structures, bearing a residual charge of (a) 0, (b) 1 or (c) 2 droplets (Q0=unit droplet charge). Because residual
charge acts repulsively to new approaching droplets, the field-focussing effect is overtaken by repulsion already for one residual droplet charge.
This illustrates that also from the charging point of view, complete vapourization between consecutive droplet ejections is a requirement for ENA
NanoDrip printing. Scale bar, 50nm.

volume. At least for a droplet of the size of the meniscus, this may
be understood by scaling the flow rate with the help of the classical
Poiseuille equation (Supplementary Methods).
Nanostructure growth initiation. The cyclic release and deposition
of spherical droplets, one at a time, in nanodripping allows enough
time for the efficient removal of solvent by vapourization during e.
This lets nanoparticles of a multitude of footprints accumulate into
a tight scaffold, which only increases in height while maintaining
constant lateral size. This mechanism is in great contrast to that of
cone-jet printing in which the ink and the accompanying nanoparticles continuously spread on the substrate owing to the high immediate volume flow rates7,8.
Fast vapourization of a droplet implies that to avoid clogging at the nozzle, the volumetric rate of vapourization at a single
deposited droplet must be higher than that at the meniscus itself
(Supplementary Methods). Our experiments show that this is fulfilled, if the ink is able to sufficiently wet the substrate surface14.
Indeed, the employed glass slips exhibited excellent wetting properties with contact angles <10. In addition, we registered very strong
wetting also when depositing optically observable volumes of solvent onto dried nanoparticle layers. This behaviour is expected due
to the solvent-specific coating of the nanoparticles. In contrast to the
outcome on wetting substrates (that is, nanopillars with homogeneous diameter), the deposition of droplets onto only partially wetted
substrates (contact angle ~60; see Methods) resulted in nano
particle scaffolds with spherical cap geometry several times larger
than a single droplet (Supplementary Fig. S7). Such a shape is clear
evidence of a single drying sessile ink droplet containing mobile
nanoparticles15 and illustrates the accumulation of several smaller
droplets on the substrate due to insufficient vapourization rates.
An interesting and important feature of ENA NanoDrip printing
is the lack of a coffee-stain effect16 and the equality of droplet and
footprint size despite strong wetting of the ink. The lack of a coffeestain pattern can be explained by the low Peclet number in the range
of unity17, giving rise to diffusion-dominated nanoparticle motion
(Supplementary Methods). However, the small size of footprints
requires further studies that go beyond our current work. We believe
that the underlying mechanism of this unexpected behaviour is that
size-related strong viscous damping causes a soft-landing process despite the high impact velocities (Supplementary Fig. S8 and
Supplementary Methods) whereas electric forces are not found to
notably enhance impact spreading (Supplementary Table S1 and
Supplementary Methods). Hence, a droplet will initially be deposited
on the substrate with a diameter similar to that of its spherical state,
and only afterwards, capillarity-driven wetting will slowly increase
its diameter18 (Supplementary Methods). The observed equality of
droplet and footprint sizes implies that, in our case, nanoparticle

settling occurs before the longer wetting phase. However, since the
time that a nanoparticle needs to diffuse along the droplet domain is
similar to the expected wetting time, diffusion alone is not expected
to allow full nanoparticle settling during the required timescale
(Supplementary Table S2 and Supplementary Methods). We speculate that because of the small size of the deposited droplets, longrange particle-substrate forces, especially Coulomb forces, extend
over the entire droplet domain, thus confining the particle settlement process17,19. The simultaneous presence of ionic charge and
external electric field renders rigorous quantification difficult and
requires further study. We note in passing that footprint patterns
did not differ for nanodroplets of opposite charge.
Structure growth by electrostatic nanodroplet autofocussing.
Once the first few droplets have laid the foundation of a pillar
(Fig. 1a), the electrostatic field induced at its strongly curved surface focusses the following incoming nanodroplets to the tip of
the forming pillar. Because of the zero lateral electric field component at the nozzle-substrate axis, droplets stay in a narrow path
even when working at nozzle-substrate distances as high as ~5m
(further increase in this separation can eventually lead to unfocussed
structure growth; see Supplementary Fig. S9). Figure 4ac show the
results of electrostatic simulations for three scenarios of different
residual ionic charge originating from the deposited charged droplets. The figures suggest that electrostatic autofocussing only acts on
droplets in close proximity to a growing nanostructure and only if
the deposited charge is sufficiently removed. Figure 4a represents
the most optimal situation of full charge removal during the course
of solvent vapourization. In this case, the metal structure clearly
acts attractive on approaching nanodroplets. Simulation results for
residual charge, equal to that carried by 1 droplet and 2 droplets,
respectively, (Fig. 4b,c, respectively) imply that, on a non-conductive
substrate liquid- and concomitant-charge accumulation would
lead to detrimental repulsion effects between pillar structure and
approaching charged droplet. Thus, observation of high aspect ratio
pillars, as that shown in Fig. 5a with 50-nm width and 850-nm
height, confirms the need for rapid liquid vapourization carrying
away ionic charge. In combination with the processes enabling uniform footprint patterns at the same size as a droplet, ENA leads to
an equality of droplet and pillar diameter. We verified this important feature by analysing the diameters of nanopillars printed at a
set of voltages and plotting them together with measured footprint
sizes (Supplementary Fig. S10).
The combination of small droplet trajectory divergence and electrostatic autofocussing allows not only a remarkable aspect ratio in
the formation of nanostructures, but also a superb spatial control of
structural placement as illustrated by an array of printed ~80-nm
wide nanodots of 1m lattice constant (Fig. 5b). The deviation

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

Height (nm)

nature communications | DOI: 10.1038/ncomms1891

40
20
0
0

100 200 300 400 500 600 700 800 900


Profile width (nm)

Figure 5 | SEM micrographs of printed nanostructures with dimensionalities from 0D to 2D. (a) Gold nanopillar of diameter ~50nm and aspect ratio
of ~17 (Scale bar, 200mm). (b) Top and (c) side view of nanopillars printed subsequently at 200nm center-to-center distance (scale bar, 200nm).
(d) 80-nm wide dots printed into a 1-m lattice constant array (1m scale bar). (e) Printed tracks with pitch sizes of 250, 200, 150, 100 and 75nm
(scale bar, 2 m). The inset shows AFM (full black lines) and SEM (red dashed lines) profiles of 150-nm pitch size. The height of AFM profiles is
given in nanometers. The SEM profiles are in arbitrary units. Tracks have reproducible heights of ~40nm and are well separated.

between the actual and aspired positions is in the range of only a few
nanometers. The spacing between structures can be further reduced
as shown by a pair of sequentially printed nanopillars of diameter 120nm, separated by a gap of 80nm (Fig. 5c,d). This example
emphasizes that moving the central impact region by an amount as
small as the structure diameter is sufficient to prevent droplets from
being attracted to the previously printed nanopillar.
The attainable positioning accuracy and resolution are further demonstrated in Fig. 5e with an example of printed in-plane
tracks. These were created by introducing a constant relative movement between the nozzle and the substrate during droplet ejection.
Figure 5e displays tracks at pitch sizes ranging from 75nm to
250nm, where full separation is achieved for track distances down
to 100nm. Atomic force microscopy profiles further reveal reproducible track heights of about 40nm (Fig. 5e, inset). Such tracks may
be employed as electrical conductors. Supplementary Figure S11
provides an example of a printed track after an annealing treatment,
proving that those structures do not loose integrity when being
sintered but instead form one entity. Quantitative results on
the electrical behaviour of printed tracks in relation to different
annealing treatments will be dedicated to future studies.
Flexibility towards growth direction and material. Interestingly,
ENA also allows fabrication of nanopillars at a variety of tilting


angles, paving the way towards the formation of a wide palette of


out-of-plane nanostructures. These tilted structures are created by
a slow lateral movement of the piezo-stage during the deposition
process. The resulting shift of the projected droplet impact position with respect to the apex of a growing nanostructure will induce
ENA to bend the trajectory of the approaching droplet towards the
nanostructure extremity. The direction at which a droplet impacts
at the structure apex depends on the velocity of the stage movement and ultimately on the velocity at which the structure grows.
For example, if the piezo-stage moves at half the structure growth
velocity, the nanostructure will grow at a 45 tilting angle. This could
be reproduced by growing tilted nanopillars at varying piezo-stage
velocities, where the tilting angle was found to be proportional to
the velocity of the substrate movement (Fig. 6).
Finally, we point out that ENA NanoDrip printing can be easily
adopted for growth from materials other than gold or even nonmetals as long as their dielectric constants are higher than that of
the surrounding medium. To demonstrate this, we have successfully
printed nanopillars made of silver and zinc oxide (Supplementary
Fig. S12).
ENA NanoDrip printing for plasmonics. ENA NanoDrip printing paves the way for easy and versatile creation of a large variety
of complex nanostructures. An application of particular interest in

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

nature communications | DOI: 10.1038/ncomms1891

1.71 m s1
34

2.95 m s1
expected 0

Figure 6 | Structures printed at varying substrate movement velocity.


Tilted nanopillars generated by ejecting nanodroplets during relative
nozzle-substrate movement at linearly increasing substrate velocities.
Pillars have a diameter of ~55nm (scale bar is 500nm). The velocities
and immediate pillar-tilting angles at the positions marked with an arrow
are displayed in the SEM micrograph. It is found that the tilting angle
is directly proportional to the employed substrate velocity. From this
linear relationship, the structure growth velocity can be deduced, which
is ~3ms1. If the substrate movement is above 3ms1, the pillars will
merge with the substrate and generate a track. That the pillar already
merges with the substrate at lower velocity is due to the non-zero pillar
diameter and the still increasing velocity, which prohibit the pillar to
elevate above the substrate. The separation between pillars is not induced
manually, by briefly interrupting ejection, but occurs spontaneously during
the process. This phenomenon may be related to the inhomogeneous
nature of the electric vector field (Supplementary Discussion).

1.0
0.8
Intensity (a.u.)

1
0.75 m s1 1.23 m s
50
65

0.6
0.4
0.2

the past half-a-decade concerns metamaterials20 and optical nano


antennas21 made of plasmonic nanostructures in various shapes
such as split ring, split-rod, bow-tie and so on. Currently, these
structures can only be fabricated using e-beam lithography (EBL) or
focussed-ion-beam (FIB) milling, both of which are very expensive
and thus not available to many researchers. ENA NanoDrip printing
promises to deliver the same performance for optical antennas and
metamaterials. As an example, we have printed a so-called Yagi-Uda
antenna (optical analogue of the old TV antennas). Figure 7a displays a SEM image of this structure. An antenna using the same
design parameters (structure size and separations) was recently realized using EBL and demonstrated a directional emission pattern22.
Having mastered the required geometrical parameters, one might
wonder whether the structures resulting from the assembly of small
colloidal gold nanoparticle have proper plasmonic response. To
investigate this, we have printed a nanopillar with a base diameter
of 50nm and height of 200nm and annealed it at 260C. In Fig. 7b,
we plot the plasmon spectra recorded under illumination polarized
perpendicular (transverse) and parallel (longitudinal) to the nanopillar long axis. Resonances at 528nm and 660nm are in reasonable
agreement with the theoretical predictions of 525nm and 700nm,
respectively23. The spectra before the annealing process deviate
from these values and are found in the Supplementary Fig. S13.
We remark that recent works have shown that, in fact, structures
as simple as single plasmonic nanospheres24, nanodisks25 or nanorods26 can act as optical antennas with strong scattering properties and large optical near-field enhancement20,21,24. Printing such
plasmonic nanostructures may also be employed for applications
in gas sensing27, graphene pholtovoltaics28 or low-energy photon
detection29.

Discussion
In this work, we have presented feature sizes down to 50nm by using
nozzle diameters in the range of 1m. Downscaling these by only
a factor of 25 would make our method fully competitive with the
state-of-the-art performances of EBL or FIB. Observed footprints,
originating from a 600-nm nozzle, have already reached ~35nm.
Besides reducing the size of employed nozzles, even smaller droplets could be achieved by pinning the meniscus not at the outer,
but at the inner opening30, for example, by treating the nozzle

0.0
500

550

600
650
700
Wavelength (nm)

750

800

Figure 7 | ENA NanoDrip printing for plasmonic applications. (a) SEM


micrograph of the ENA NanoDrip-printed geometry of a recently demons
trated Yagi Uda antenna22 (scale bar, 200nm). (b) The measured scat
tering spectrum of a printed and subsequently annealed gold nanopillar
with a diameter of 50nm and aspect ratio of ~4. The graph shows spectra
for longitudinal (black line) and transverse (red line) excitation. The small
peak at 660nm for transverse polarization is due to slight coupling of the
excitation into the longitudinal mode.

surface with a solvent-repelling coating1,30. This may allow droplets as small as 20nm from a 600nm nozzle at a manageable effort.
Further downscaling may be achieved by employing ever smaller
nozzles. Along this line, we believe that a lower limit to the process
will mainly be induced by difficulties of further nozzle downscaling
(also the involved clogging issues) or size restrictions imposed by
the nanoparticles (as particles should still be substantially smaller
than the structures they build). Of great importance for any technical advancement will also be a better understanding of the different
physical aspects. We have identified several stringent conditions,
which have to be fulfilled for proper functioning of ENA NanoDrip
printing. Most important is a proper adjustment of the diverse flow
rates, namely of ejection flow rate and vapourization flow rate at
both, droplet and meniscus (see also the Supplementary Methods,
concerning this topic). The implementation of strongly deviating
solvents (for example, with respect to surface tension, viscosity,
vapour pressure and so on) will not be successful at the same conditions used here. In this respect, it will be of importance to also
include external factors that we have not yet investigated (for example, temperature). We believe that both downscaling and further
process versatility will be accompanied with a better understanding of the process. In contrast to techniques like EBL and FIB, ENA
NanoDrip printing is orders of magnitude less expensive, does not
require vacuum, and can easily create out-of-plane structures. We
anticipate that availability of reproducible microfabricated print

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

nature communications | DOI: 10.1038/ncomms1891

heads31 in the future will introduce further controllability and parallelization of the process with a multitude of nozzles, providing an
on-demand, bottom-up and scalable nanofabrication technique.

diameter D). The charge relaxation time describes the process of relaxing charges
on a conductor surface, that is, the time it takes for the surface potential to settle
after a voltage stimulus and is defined as

te =

Methods

Printing process. The set-up shown in Fig. 1a consists of an in-house built micro
scope equipped with a 3D piezo-stage (MadCityLabs), electrical equipment for
pulse generation and a central control unit. A glass substrate was placed on top
of an ITO coated glass slide mounted at the piezo-stage. The piezo-stage had a
working distance of 300m. In the respective area, structures can be well aligned
at the inherent precision of the piezo-stage (generally <10nm). For optical detection, this microscope worked in the iSCAT mode32. This allowed highly sensitive
detection of printed nanostructures and tight focussing (low view of depth). The
latter was essential for controlling the nozzle-substrate distance (see below for
more information). For the respective experiments, the wettability of the substrate
with respect to the Au ink was reduced by vapour phase coating of a self-assembled
monolayer of 1H,1H,2H,2H-perfluorooctyltrichlorosilane, resulting in a contact
angle of ~60.
Glass pipettes with outer diameters between 550nm and 1,300nm were
prepared with a Sutter P-97 pipette puller. Only pipettes with good nozzle quality,
that is, having no cracks, chipped edges or the like, were used. Ink was introduced
to the backside opening of the pipette and drawn to the very tip by capillary forces
without any applied pressure. The outer pipette wall and a small region of the
nozzle tip interior were coated with a 10-nm adhesion layer of titanium followed
by 100nm of gold using e-beam evaporation. Irrespective of the small nozzle size,
only minimal caution was necessary when handling them. The nozzle-substrate
distance was assessed with a precision of about 150nm by tightly focussing at the
capillary tip and then moving the focal plane at a controlled distance by a movement of the micro-stage holding the objective. The substrate was then placed into
focus by a z-movement of the piezo-stage. For most experiments, the employed
ink contained 37nm gold particles in cyclododecene (ULVAC Technologies). The
ink was diluted by a factor of 50 with n-tetradecane to arrive at a final concentration of ~0.1 vol%. Dilution was necessary to avoid clogging, at least for a period of
several minutes. At substantially higher concentrations, rapid clogging prohibited
a smooth workflow. Early clogging could normally be reversed by the application
of a high-voltage pulse (up to 400V). For calculations, the physical properties
of n-tetradecane were adopted. The measured conductivity of the diluted ink
was 6.8108Sm1. Silver (ULVAC Technologies) and zinc oxide (Nanograde)
inks were both based on n-tetradecane and contained nanoparticles with sizes of
37nm and ~25nm, respectively. Voltage signals were generated by a waveform
generator (RIGOL) and amplified by a homemade HV amplifier to up to 400Volts.
Annealing and optical measurements. Thermal annealing was performed in a
rapid thermal annealing oven (Jipelec). During purge-pumping by nitrogen gas,
the temperature was first ramped at 2Cs1 to 100C, then at 10Cs1 to 200C
and finally at 3Cs1 to 260C. Once the temperature reached 260C, the heat
source was deactivated and a 5% mixture of H2/N2 was introduced at 500s.c.c.m.
The sample was passively cooled to room temperature. Scattering spectra were
obtained in a low grazing angle arrangement (Supplementary Fig. S14).
Requirements for the nanodripping mode. To achieve the nanodripping mode,
the characteristic time of supply of liquid to the droplet (q) has to be longer than
the characteristic time of drop formation11. Here q is estimated as the duration
between consecutive droplet ejections (e) and therefore equals the droplet volume
(~d3) divided by the volume flow rate V .

tq = te =

pd3
6V

(1)

The characteristic time of drop formation d is based on the intrinsic properties of


the fluid with denoting the liquid viscosity and the liquid surface tension11

td =

md
g

(2)

The drop formation time states how long it takes for a liquid element, ejected
through a circular exit with diameter d, to form a droplet solely by the action of
surface tension and viscous resistance. If the characteristic time of supply of liquid
is shorter than the drop formation time, the ensuing convective liquid flow will not
allow for the relaxation of the liquid element into a droplet. Instead, a continuous
jet can form, which would be the basis of the cone-jet ejection mode.
A timeframe right before droplet ejection is schematically depicted in Fig. 2c,
where the meniscus diameter is assumed to be equal to the outer nozzle diameter.
During ejection, the meniscus has to be quasi-steady, which requires the charge
relaxation time to be smaller than the time of supply of liquid to the meniscus
Q (the latter is also given by equation (1) but with d replaced by the meniscus


e
s

(3)

Here and are the liquid electric permittivity and conductivity. If the charge
relaxation time is longer than the time of supply of liquid, more charge would
be removed from the surface by liquid flow than what is delivered by bulk
electrical conduction. The surface charge would therefore diminish causing the
meniscus to inherently collapse. In Supplementary Fig. S15 a distribution of
these important timescales is shown, proving that the requirements leading
to nanodripping ejection are fulfilled for the entire regime investigated in
this study.
Numerical simulations. Numerical simulations for the determination of the
electric field were performed by COMSOL multiphysics software in the AC/DC
electrostatics module, using the linear UMFPACK solver. The geometry of the
pipette was determined from SEM micrographs and modelled as axisymmetric.
The model of the nozzle region includes the pipette, a pulled liquid meniscus and
a small pendant droplet at its tip, all of which are assumed to be at equipotential.
The grid was automatically refined until results deviated by less than 2%. To
estimate the droplet diameter, d, the following balance of surface tension and
electric forces has to be fulfilled:

e0
E 2 (d ,V ) daz = p dg
2

(4)

Here represents the surface tension, E is the electric field, the permittivity of air
and daz a differential surface element for the integration over the small hemispherical surface with subscript z denoting the z-direction. A surface plot of the obtained
electric field can be found in Fig. 2c. The integral in equation (4) was evaluated by
solving for the Maxwell stress tensor in the z-direction followed by integration on
the surface of the small pendant droplet. Input voltages were iteratively adjusted
until the resulting electric force matched the surface tension force with an accuracy
of better than 2%.

References

1. Park, J. U. et al. High-resolution electrohydrodynamic jet printing. Nat. Mater.


6, 782789 (2007).
2. Ahn, B. Y. et al. Omnidirectional printing of flexible, stretchable, and spanning
silver microelectrodes. Science 323, 15901593 (2009).
3. Schirmer, N. C., Schwamb, T., Burg, B. R., Hotz, N. & Poulikakos, D.
Controlled free-form fabrication of nanowires by dielectrophoretic dispension
of colloids. Appl. Phys. Lett. 95, 033111 (2009).
4. Fuller, S. B., Wilhelm, E. J. & Jacobson, J. M. Ink-jet printed nanoparticle
microelectromechanical systems. J. Microelectromech. Syst. 11, 5460
(2002).
5. Byun, D. et al. Pole-type ground electrode in nozzle for electrostatic field
induced drop-on-demand inkjet head. Sensor Actuat. A-Phys. 141, 506514
(2008).
6. Velev, O. D. & Gupta, S. Materials fabricated by micro- and nanoparticle
assembly - the challenging path from science to engineering. Adv. Mater. 21,
18971905 (2009).
7. Mishra, S., Barton, K. L., Alleyne, A. G., Ferreira, P. M. & Rogers, J. A. Highspeed and drop-on-demand printing with a pulsed electrohydrodynamic jet.
J. Micromech. Microeng. 20, 095026 (2010).
8. Wang, K. & Stark, J. P. W. Deposition of colloidal gold nanoparticles by fully
pulsed-voltage-controlled electrohydrodynamic atomisation. J. Nanopart. Res.
12, 707711 (2010).
9. Li, D. & Xia, Y. N. Electrospinning of nanofibers: Reinventing the wheel?
Adv. Mater. 16, 11511170 (2004).
10. GananCalvo, A. M., Davila, J. & Barrero, A. Current and droplet size in the
electrospraying of liquids. Scaling laws. J. Aerosol. Sci. 28, 249275 (1997).
11. Jaworek, A. & Krupa, A. Classification of the modes of EHD spraying.
J. Aerosol. Sci. 30, 873893 (1999).
12. Cloupeau, M. & Prunetfoch, B. Electrohydrodynamic spraying functioning
modes - a critical-review. J. Aerosol. Sci. 25, 10211036 (1994).
13. Carson, R. S. & Hendrick, C. D. Natural pulsations in electrical spraying of
liquids. AIAA J. 3, 10721075 (1965).
14. Hu, H. & Larson, R. G. Evaporation of a sessile droplet on a substrate. J. Phys.
Chem. B 106, 13341344 (2002).
15. Kuncicky, D. M. & Velev, O. D. Surface-guided templating of particle assemblies
inside drying sessile droplets. Langmuir 24, 13711380 (2008).
16. Deegan, R. D. et al. Capillary flow as the cause of ring stains from dried liquid
drops. Nature 389, 827829 (1997).

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

ARTICLE

nature communications | DOI: 10.1038/ncomms1891


17. Widjaja, E. & Harris, M. T. Particle deposition study during sessile drop
evaporation. AIChE J. 54, 22502260 (2008).
18. Dong, H. M., Carr, W. W., Bucknall, D. G. & Morris, J. F. Temporally-resolved
inkjet drop impaction on surfaces. AIChE J. 53, 26062617 (2007).
19. Bhardwaj, R., Fang, X. H., Somasundaran, P. & Attinger, D. Self-assembly of
colloidal particles from evaporating droplets: role of DLVO interactions and
proposition of a phase diagram. Langmuir 26, 78337842 (2010).
20. Ozbay, E. Plasmonics: merging photonics and electronics at nanoscale
dimensions. Science 311, 189193 (2006).
21. Novotny, L. & van Hulst, N. Antennas for light. Nat. Photon 5, 8390 (2011).
22. Curto, A. G. et al. Unidirectional emission of a quantum dot coupled to a
nanoantenna. Science 329, 930933 (2010).
23. Moroz, A. Depolarization field of spheroidal particles. J. Opt. Soc. Am. B 26,
517527 (2009).
24. Kuhn, S., Hakanson, U., Rogobete, L. & Sandoghdar, V. Enhancement of singlemolecule fluorescence using a gold nanoparticle as an optical nanoantenna.
Phys. Rev. Lett. 97, 017402 (2006).
25. Lippitz, M. et al. Nanoantenna-enhanced ultrafast nonlinear spectroscopy of a
single gold nanoparticle. Nat. Commun. 2, 333 (2011).
26. Mohammadi, A., Sandoghdar, V. & Agio, M. Gold nanorods and nanospheroids
for enhancing spontaneous emission. New J. Phys. 10, 105015 (2008).
27. Liu, N., Tang, M. L., Hentschel, M., Giessen, H. & Alivisatos, A. P.
Nanoantenna-enhanced gas sensing in a single tailored nanofocus. Nat. Mater.
10, 631636 (2011).
28. Echtermeyer, T. J. et al. Strong plasmonic enhancement of photovoltage in
graphene. Nat. Commun. 2, 458 (2011).
29. Knight, M. W., Sobhani, H., Nordlander, P. & Halas, N. J. Photodetection with
Active Optical Antennas. Science 332, 702704 (2011).
30. Stachewicz, U., Dijksrnan, J. F., Burdinski, D., Yurteri, C. U. & Marijnissen, J. C.
M. Relaxation times in single event electrospraying controlled by nozzle front
surface modification. Langmuir 25, 25402549 (2009).
31. Lee, J. S. et al. Design and evaluation of a silicon based multi-nozzle for
addressable jetting using a controlled flow rate in electrohydrodynamic jet
printing. Appl. Phys. Lett. 93, 243114 (2008).

32. Lindfors, K., Kalkbrenner, T., Stoller, P. & Sandoghdar, V. Detection and
spectroscopy of gold nanoparticles using supercontinuum white light confocal
microscopy. Phys. Rev. Lett. 93, 037401 (2004).

Acknowledgements

Financial support for the work reported in this paper provided by the Swiss National
Science Foundation through grant number 2-77485-09 is gratefully acknowledged. The
authors would also like to thank Dr. M. Tiwari of the Laboratory of Thermodynamics in
Emerging Technologies of ETH Zurich and Dr. Mario Agio of the Nano-Optics group of
ETH Zurich for useful discussions.

Author contributions

P.G. and J.S. built the set-up, designed and analysed experiments. P.G. interpreted data
and performed FEM simulations. P.G. and S.K. designed the model for the simulation
and performed experiments. P.G., D.P. and V.S. wrote the manuscript, which was edited
by all authors. H.E. designed and performed the optical measurements and interpreted
optical spectra. D.P. supervised all aspects of the project, designed the research and gave
scientific and conceptual advice. V.S. supervised the optical measurements and gave
scientific advice.

Additional information

Supplementary Information accompanies this paper at http://www.nature.com/


naturecommunications
Competing financial interests: The authors declare no competing financial interests.
Reprints and permission information is available online at http://npg.nature.com/
reprintsandpermissions/
How to cite this article: Galliker, P. et al. Direct printing of nanostructures
by electrostatic autofocussing of ink nanodroplets. Nat. Commun. 3:890
doi: 10.1038/ncomms1891 (2012).

nature communications | 3:890 | DOI: 10.1038/ncomms1891 | www.nature.com/naturecommunications

2012 Macmillan Publishers Limited. All rights reserved.

Vous aimerez peut-être aussi